You are on page 1of 9

Available online at www.sciencedirect.

com

ScienceDirect
Acta Materialia 78 (2014) 1422
www.elsevier.com/locate/actamat

Three-dimensional imaging of hydrogen blister in iron


with neutron tomography
Axel Griesche a,, Eitan Dabah a, Thomas Kannengiesser a, Nikolay Kardjilov b,
Andre Hilger b, Ingo Manke b
a

BAM Federal Institute for Material Research and Testing, Department 9 Component Safety, Unter den Eichen 87, 12205 Berlin, Germany
b
Helmholtz-Zentrum Berlin (HZB), Institute of Applied Materials, Hahn-Meitner-Platz 1, 14109 Berlin, Germany
Received 7 April 2014; received in revised form 13 June 2014; accepted 15 June 2014
Available online 10 July 2014

Abstract
We investigated hydrogen embrittlement and blistering in electrochemically hydrogen-charged technical iron samples at room temperature. Hydrogen-stimulated cracks and blisters and the corresponding hydrogen distributions were observed by neutron tomography.
Cold neutrons were provided by the research reactor BER II to picture the sample with a spatial resolution in the reconstructed threedimensional model of 25 lm. We made the unique observation that cracks were lled with molecular hydrogen and that cracks were
surrounded by a 50 lm wide zone with a high hydrogen concentration. The zone contains up to ten times more hydrogen than the bulk
material. The hydrogen enriched zone can be ascribed to a region of increased local defect density. Hydrogen also accumulated at the
sample surface having the highest concentration at blistered areas. The surfaces of the brittle fractured cracks showed micropores visualized by scanning electron microscopy. The micropores were located at grain boundaries and were surrounded by stress elds detected
by electron backscattered diraction. The cracks clearly originated from the micropores.
2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Hydrogen embrittlement; Hydrogen diusion; In situ; Neutron tomography; Iron

1. Introduction
The durability of alloys can be inuenced signicantly
by hydrogen uptake, leading to a degradation of the
mechanical properties with possible subsequent hydrogenassisted cracking (HAC) [1,2]. In numerous investigations,
dierent factors aecting the hydrogen embrittlement have
been studied, e.g. hydrogen content [3], residual stress and
strain due to internal or external forces [4] and microstructure [5]. All these factors can lead to a critical condition for
hydrogen embrittlement through specic mechanisms. The
most popular are called hydrogen-enhanced decohesion

Corresponding author. Tel.: +49 30 8104 3990; fax: +49 30 8104 1557.

(HEDE) and hydrogen-enhanced localized plasticity


(HELP) [6,7].
The uptake of hydrogen into the material can occur via
dierent ways. One way for example is the diusion of dissociated hydrogen from the plasma of the protection gas
during arc welding into the weld pool [8] where it remains
after solidication. Another way could be the diusion from
the surface into the bulk due to corrosion processes [9].
However, the materials degradation process is enhanced
drastically if hydrogen is introduced rapidly exceeding
locally the solubility limit. Then, recombination of hydrogen may yield hydrogen-assisted formation and growth of
microvoids. Subsequently, HAC, also referred to as cold
cracking, can cause failure of the material or component,
respectively.

E-mail address: axel.griesche@bam.de (A. Griesche).


http://dx.doi.org/10.1016/j.actamat.2014.06.034
1359-6454/ 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

A. Griesche et al. / Acta Materialia 78 (2014) 1422

Another obvious way of obtaining cavities is by casting


a metal containing a large amount of dissolved gases. In
this case the solubility in the melt is larger than in the solid.
However, we are not addressing the casting and solidication problems here. The emphasis will be placed on the formation and hydrogen assisted growth of cavities in the
solid state.
In the past, various investigations, summarized by Condon and Schober [10], treated the problem of pores due to
hydrogen and especially of subsurface pores. The latter
produced so-called blisters that are visible on the surface
as bulges. They discussed three basic cavity formation
mechanisms with near-surface plastic deformation as the
one responsible for blistering. Ravi et al. [11] focused on
this specic mechanism in their investigation on microalloyed steels. They found hydrogen-induced blister cracking (HIBC) when steel is exposed to moist H2S. The H2S
can dissociate at the surface and releases atomic hydrogen
to the interior. This atomic hydrogen in turn can be
trapped by subsurface precipitates or defects and accumulates to form blisters lled with molecular hydrogen. More
recently, Escobar et al. [12] investigated the inuence of
various hydrogen charging parameters on blistering,
namely the eect of dierent electrolytes, current densities
and charging times in pure iron and in ferritebainite steels.
They found that in acid electrolytes with high current densities, blister occurred more easily in pure iron than in steel.
Others detected blistering e.g. in Mg alloys [13], Ti alloys
[14], W [15], Al and Mo [16] and Si [17].
The experimental evidence of the models explaining
HIBC is often proven indirectly because methods to detect
hydrogen directly are rare. Such direct and mostly nondestructive techniques have been already used to measure
hydrogen in metals. Hoshihira et al. [18] examined the
hydrogen distribution around blisters by tritium autoradiography (TARG) and in the case of blisters in Mo [16] by
a tritium imaging plate technique (TIP). Tritium decays
and emits b-electrons that expose a photographic lm on
the samples surface. After exposure, the photographic lm
was developed and xed to obtain Ag precipitates with a
diameter of 0.3 lm. The two-dimensional (2-D) distribution of Ag precipitation was observed by SEM and energy
dispersive X-ray spectroscopy (EDS) and corresponds to
the tritium distribution in the samples surface. Hanada
et al. [19] used TARG to measure local hydrogen concentration proles around imperfections in tempered martensitic
steel and Katano et al. [20] used high-strength steels. Garet
et al. [21] applied TARG to visualize tritium located on nonmetallic inclusions in CrMo low alloy steels. Although concentrations as small as 0.1 wt. ppm could be detected, the
time resolution was in the order of some 10 min only.
Revay et al. [22] used prompt c activation analysis
(PGAA) to measure hydrogen in catalytic reactors at concentrations in the wt. ppm range but with limited spatial
(0.2 mm) and temporal (hours) resolution. PGAA is
based on the radiative neutron capture of nuclei, i.e. the
(n, c) reaction. Each isotope of every chemical element

15

(except 4He) is capable of absorbing a neutron, followed


by an immediate release of the binding energy in the form
of c radiation.
Neutron radiography (NR) is a modern non-destructive
imaging method with a wide eld of applications in materials science, physics, biology and archaeology (see Refs.
[23,24] for an overview). Neutrons in research facilities
are produced either by a chain reaction of nuclear ssion
in a nuclear reactor or by nuclear spallation. The latter process involves a particle accelerator, pounding the particles
on a heavy metal target to expel a beam of neutrons. Special NR variants are e.g. neutron tomography (NT) with
polarized neutrons in order to investigate magnetic elds
in three dimensions [25,26] or imaging with scattered neutrons [2729]. The progress in detector techniques and the
increase in brilliance and ux of neutron sources mean that
fast dynamic processes like hydrogen eusion in technical
iron can now be measured quantitatively [30].
In this work, we used NT to measure the three-dimensional (3-D) hydrogen distribution in technical iron. We
used electrochemically charged samples, which well mimics
hydrogen deposition from applications such as arc welding
or from corrosion. In order to simplify the investigation of
HAC mechanisms and especially the blistering eect we
have chosen technical iron (ARMCOe) as the material
because it has a low solubility for hydrogen, consists of a
single phase and has only a few hydrogen trapping site
types. Thus, complex eects originating from a multiphase
microstructure, phase boundaries and inclusions as well as
from alloying elements or from hydrogen-induced phase
transformations can be excluded.
The aim of this study is to reveal how electrochemically
charged ferrite forms hydrogen-induced blister cracks. The
novelty of this research is the use of a highly brilliant neutron beam at the new neutron radiography instrument V7
at Helmholtz-Zentrum Berlin (HZB) with the newest detector generation that allows the detection and visualization
of hydrogen distributions with an unprecedented spatial
resolution. An extended fracture analysis adds information, which is necessary to understand the crack initiation
and blister evolution.
2. Experimental
2.1. Sample characterization
The used sample material was ARMCOe iron supplied
as rolled plates. The microstructure of the technical iron
entails solely ferritic a-phase and is composed of grains
with the average size of 80 lm, as shown in Fig. 1. Inclusions or imperfections other than grain boundaries were
not visible by optical microscope inspection.
The chemical composition of the material is given in
Table 1.
For the NT experiments, small coupons were cut from
the plates in the as-delivered state with dimensions
45 mm height  10 mm width  5 mm thickness. This

16

A. Griesche et al. / Acta Materialia 78 (2014) 1422

Fig. 1. Light microscope image of a polished and etched (2% HNO3 Nital
solution) ARMCOe iron sample cross-section. A predominant orientation of the grains with respect to the rolling direction is not visible.

Table 1
Chemical composition of the iron material (in wt.%).
Fe

Mn

Ni

Cr

Mo

>99.7

0.002

0.05

0.002

0.013

0.013

0.002

0.003

geometry allows fast hydrogen charging due to the large


surface to volume ratio and guarantees the development
of blister when using standard charging parameters. This
sample geometry is also suitable for neutron tomography.
The coupons surfaces were cleaned by sandblasting in
order to remove any residuals prior to the cathodic hydrogen charging. The samples were electrochemically charged
with a galvanostatic wiring in an aqueous 0.5 M H2SO4
solution containing 0.25 g l 1 sodium arsenide as surface
recombination inhibitor for hydrogen. The current density
was 50 A m 2. The samples were charged for 36 h, a sucient time for the formation of blisters already visible by
the eye on the samples surface during the charging process.
The hydrogen content introduced in the sample was measured immediately (<10 min) after the electrolytic charging
process by carrier gas hot extraction (CGHE). The hydrogen concentration of a coupon was 0.6 wt. ppm determined
by heating the sample freely in the pre-heated furnace of the
CGHE analyzer (Bruker JUWE H-MAT221) from room
temperature up to 1173 K. The concentration value is
gained by integrating the sensors signal over time and normalizing to the samples weight. About 80% of the hydrogen, i.e. 0.5 wt. ppm, is already collected after 5 min. We
get a similar result when both furnace and sample are
heated rapidly from room temperature up to 673 K with a
subsequent isothermal annealing phase. Here 90% of the
hydrogen desorbed from the sample within 5 min.

blisters, we used a neutron tomographic method. Neutrons


have a strong interaction with the proton of the hydrogens
core, which is signicantly higher than the interaction of
neutrons with the isotopes of iron. This allows imaging
with clear contrast between hydrogen-enriched and hydrogen-depleted regions. Although hydrogen concentrations
as low as 20 wt. ppm can be detected in iron [31], such
low concentrations have not yet been measured in three
dimensions with a spatial resolution on the lm scale. The
neutron tomography measurements were performed at
the BER II research reactor [32,33] at HZB.
We used a new high resolution detector system to obtain
spatially resolved transmission data of the hydrogen
charged technical iron samples. A Gd2O2S scintillator with
10 lm thickness detected the transmitted neutrons. The
light produced by the scintillator was transferred via a
macro lens optic to a highly sensitive CCD camera (Andor
DW436N-BV with 16-bit grey value resolution). The pixel
size is 6.4 lm. The currently possible spatial resolution of
NT is 2030 lm. The details of the set-up are described
by Williams et al. [34].
For tomographic measurements the sample was rotated
stepwise through 360 (see Fig. 2) to radiograph the sample
in dierent orientations. Overall, 600 single radiographic
projections were taken with an exposure time of 80 s
for each image. The radiograms were normalized with at
eld images (taken without the sample) and dark eld
images (taken with the neutron beam switched o to correct for, e.g. the dark current of the camera). After the measurements, the projections were processed by tomographic
reconstruction algorithms (Octopus Software) using the ltered back projection method in order to obtain a virtual
full scale 3-D model of the sample with blisters and the
hydrogen distributions in the sample.
The hydrogen charged samples with blisters were stored
in liquid nitrogen until the start of the neutron tomography

2.2. Neutron tomography


For the examination of the hydrogen distribution in the
sample and in order to observe details about cracks and

Fig. 2. Principal sketch of the tomography set-up.

A. Griesche et al. / Acta Materialia 78 (2014) 1422

to avoid eusion of the diusible hydrogen. Diusible


hydrogen is usually dened as the amount of hydrogen that
diuses out of the sample at room temperature or at moderate temperatures below 573 K. Just before mounting the
sample on the holder for the neutron tomography, it was
taken out of the freezing agent and was cleaned by rinsing
with ethanol and acetone to get rid of condensated humidity. A possible eusion of hydrogen during the tomography, which lasted for 14 h, was not detected.
3. Results
3.1. Neutron tomography
The reconstruction of a set of radiographic projections
yields a 3-D model of the sample. This allows for the rst
time the precise 3-D visualization and quantitative analysis
of hydrogen distributions in a bulk iron sample. Fig. 3
shows a part of the sample in a 3-D reconstruction.
Fig. 4 shows three slices taken out of the reconstructed
3-D model in Fig. 3 for a more detailed view. The slices
are orthogonal xy-, yz- and xz-planes with respect to the
Cartesian coordinate system presented in Fig. 2. The
images are smoothed in three dimensions by a 5  5  5
median lter to reduce noise that is generated by the reconstruction algorithm.
Three observations stand out in the slices shown in
Fig. 4. First, we observed numerous cracks in the subsurface area below the blisters and in the bulk, often aligned
in the predominant z-direction. The alignment correlates
with the rolling direction of the used plate material from
which the sample coupons were cut.
Second, we found bright regions surrounding the cracks,
which are located mostly far from the samples surface
deep in the volume of the sample. Third, the samples
surface is decorated with a thin bright zone, which is more
intense at blisters.
In order to verify that the bright areas around the cracks
are an evident proof of hydrogen presence and no scattering artefacts, we applied a standard annealing procedure

17

for removing diusible hydrogen without changing the


microstructure. Therefore, the sample was heat-treated
for 48 h at 323 K according to Ref. [35] for removal of diffusible hydrogen. Subsequently, a second tomography was
taken with the heat treated sample. The results of both
tomographies, before and after the heat treatment, are presented in Fig. 5a and b. The calculation of the hydrogen
density is described further below and the result is shown
in Fig. 5c.
The success of the heat treatment was additionally controlled with a reference sample by means of CGHE. For
this purpose, the reference sample was heated freely in a
pre-heated furnace from room temperature up to 1173 K
and the hydrogen concentration desorbing from the sample
accumulated to 0.05 wt. ppm. Obviously, the removal procedure reduced the hydrogen concentration by an order of
magnitude. The residual 0.05 wt. ppm hydrogen produces a
too low contrast in the transmission images and is thus not
visible. The remaining hydrogen is either energetically
more strongly trapped than the diusible hydrogen, e.g.
at hydrogen-induced dislocation twins, or is still present
as molecular hydrogen in micro voids. In summary, we
can exclude that the bright areas are image artifacts due
to scattering eects of the neutrons, e.g. at crack surfaces.
For the rst time, the hydrogen distribution around cracks
has been experimentally visualized.
In order to quantify the hydrogen amount in the sample,
the two tomographic volumes of the hydrogen-charged and
hydrogen-discharged sample were precisely registered and
subtracted. The values of the voxels in the resulting volume
were converted from gray levels into linear attenuation
coecients R of the residual hydrogen. The 3-D distribution of the hydrogen density q can be obtained by using
the formulae q = RuH/NA rtotal, with uH being the atomic
mass, NA the Avogadro constant and rtotal the total
cross-section of hydrogen for cold neutrons, which is
experimentally determined for the imaging beam line
CONRAD with 110 barn.
Line proles of the attenuation coecient and the
hydrogen density above a crack are plotted in Fig. 6. The

Fig. 3. Inclined view on the reconstructed 3-D model of a hydrogen-charged iron sample. The surface with blisters is shown in (a). The crack distribution
in the interior is presented in (b) and the additional hydrogen distribution in (c). Most of the cracks are lled with hydrogen except some of the cracks
underneath blisters. Sample depth is 5 mm.

18

A. Griesche et al. / Acta Materialia 78 (2014) 1422

Fig. 4. Three orthogonal slices taken from arbitrarily positions in the reconstructed 3-D model: (a) xy-plane, (b) yz-plane, (c) xz-plane. The white color
decorating internal cracks indicates the presence of hydrogen in the sample. Black areas are cracks without hydrogen. Hydrogen at the samples surface is
visible as a thin white line, which is intensied above blisters.

Fig. 5. Slices (xy-planes) of two tomographies of the same sample (a) after hydrogen charging and (b) after heat treatment. (c) Hydrogen number density
distribution of (a).

width of the crack is estimated to be 140 lm and the hydrogen seam on each side of the crack is 4050 lm wide,
which is well above the spatial resolution of the tomography method.
The dierence in the intensity minima of both curves in
Fig. 6b in the middle of the crack implies that hydrogen is
located not only at the crack surface but also in the adjacent volume. Since the intensity levels of both proles are
similar for distances larger than approximately 100 lm
away from the crack, we conclude that most of the hydrogen is located in and around cracks.
A more detailed image analysis allows for further quantication of the results. The hydrogen density in the cracks
cavity can be converted e.g. into a gas pressure assuming
an ideal gas behavior. The analysis of a set of cracks
distributed in the whole sample gained a maximum value
of the intensity peaks of 100200 wt. ppm. The intensity

distribution clearly shows that hydrogen is mostly located


at the crack surfaces and the adjacent lattice volume with
a mean width of 50 lm (see Fig. 6b). The concentration
of hydrogen gas in the cracks volume corresponds to a
pressure of 515 MPa (in Fig. 6 10 MPa), which is an
order of magnitude smaller than the yield strength of
200 MPa for non-hydrogenous rolled iron.
3.2. Blister morphology and fractography
In order to understand both the origin of blisters and the
measured corresponding hydrogen distribution, the cracks
surface morphology was characterized by SEM examination. Fig. 7a shows a cross-section polish of a blister with
corresponding cracks.
The SEM image of a view into a crack (Fig. 7b) shows a
brittle fractured surface, which is characterized by cleavage

A. Griesche et al. / Acta Materialia 78 (2014) 1422

19

Fig. 6. (a) Prole of the hydrogen number density and (b) prole of the attenuation coecient across a crack taken from (c) the dashed line in the
tomographic image. The solid line in (b) denotes the attenuation coecient prole across the same crack after the samples heat treatment.

despite the ductile characteristics of the material [36]. This


fracture mode is also common for ferritic steels and high
strength steels containing hydrogen [37]. All observed
cracks and crack surfaces have the same appearance independent of their location in the sample. It is interesting
to note is that small pores with diameters of 1 lm are
visible all over the crack surfaces (see Figs. 7b and 8). Such
pores are not visible in the uncharged samples.
We applied electron backscatter diraction (EBSD) in
order to better understand the pores origin and to get
information about the early stages of the evolution of
cracks. Fig. 9 presents small cracks, which evidently originated from a pore.
Fig. 9 snapshots the early stage of crack initiation
caused by preceding pore formation. Such pore formation

with subsequent cracking starting from the pores inner


surface has been already reported by Ren et al. [38]. The
authors assume that cracks initiate as the buildup of gas
pressure due to the recombination of hydrogen atoms to
H2 molecules exceeding the tensile strength of the hydrogenous material, which is lower than that of the hydrogenfree material. The cracks volume increases until the gas
pressure decreases below the tensile strength. Fig. 9b shows
that the pore is located on a grain boundary. Grain
boundaries are strong traps for hydrogen [39], which can
drive the accumulation of hydrogen with subsequent
hydrogen recombination and pore formation. This scenario
is supported by lattice distortions visible as dierent blue
shades of the color code for dierent ferrite orientations
in the vicinity of the original pore.

Fig. 7. SEM images of a blister in iron. The polished cross-section (a) shows cracks lifting the material up to form a bulge at the surface and the
enlargement of the white marked region presents a view into a crack (b). The white arrows mark a collection of small pores at the cracks surface.

20

A. Griesche et al. / Acta Materialia 78 (2014) 1422

4.1. Hydrogen-assisted cracking

Fig. 8. SEM image of a crack surface showing pores with diameters on the
lm scale and below.

4. Discussion
Carrier gas hot extraction up to 1173 K indicated that
after hydrogen charging the hydrogen present in the material can be classied as diusible hydrogen. Diusible
hydrogen is dened as hydrogen that desorbs below
573 K, i.e. the binding energy between trap and H atom
is DEB < 36 kJ mol 1, respectively [40]. This includes
hydrogen on interstitial sites of the metal lattice and the
reversible trapped hydrogen. The heat treatment procedure described in Section 3.1 evidently proves that the
observed hydrogen in the material is diusible hydrogen.
The mechanism of simultaneous removal of the molecular
hydrogen in the cracks remains unclear. The diusible
hydrogen has been held responsible in the literature for
hydrogen embrittlement, e.g. in high strength steels [41],
and most likely plays an important role in the blister formation process under evaluation in this work. Not
detected by CGHE, unless the sample is heated past the
melting point by melt extraction, is molecular hydrogen
from pores and cracks. Hence direct measurement methods with neutrons are advantageous because they allow
for an in situ detection of all hydrogen [42]. In the following discussion we will highlight the observed hydrogen
distributions and blisters.

We charged the samples for 36 h with a current density


of 50 A m 2 introducing much more atomic hydrogen than
the body centered cubic lattice of a-ferrite can dissolve in
thermodynamic equilibrium. The solubility limit is
reported by Choo et al. [43] to be 0.9 wt. ppm. The excess
hydrogen is located in traps although the variety of trap
sites in ARMCOe iron is limited to (in the order of trapping strength): some solute atoms, free surfaces, vacancies,
dislocations, strain elds and grain boundaries. However,
Kumnick and Johnson [44,45] reported for ARMCOe that
room temperature deformation substantially increases the
amount of trapped hydrogen by increasing the trap density
from 1020 m 3 for annealed iron to a little over 1023 m 3
for heavily deformed iron. However, since blistering was
observed already during charging, HAC with all observed
phenomena happened in parallel. We assume the establishment of a steady-state situation in which the same amount
of atomic hydrogen entering the sample is recombining and
leaving the sample via crack channels as molecules. A certain portion will diuse back to the surface and desorb
from the surface after recombination. Also after the end
of charging, excess hydrogen is present leading to HE
and subsequent HIBC. This explains the observed lm-sized
pores and the brittle fracturing (see Section 3.1).
4.2. Hydrogen accumulations
4.2.1. Hydrogen on the sample surface
Fig. 4 shows for the rst time the generally supposed
hydrogen concentration accumulation at the sample surface. Since we found these accumulations on both the
rolled and cut surfaces of the coupons, we assign clearly
this nding to the hydrogen charging process. The local
enrichment is a residual of the concentration gradient
pointing inwards during the electrochemical charging process in combination with a possible higher trap density due
to the sand blasting. Hadam and Zakroczymski [46] investigated such regions with considerably higher hydrogen
concentration near the sample surface and attributed it to
a locally increased number of traps. The estimated thickness of this region was dependent on the carbon content

Fig. 9. SEM image (a) and corresponding EBSD map (b) of cracks emanating from a pore.

A. Griesche et al. / Acta Materialia 78 (2014) 1422

of the iron alloys they investigated and was in iron


(0.05 wt.% C) 440 lm and in high-carbon steel (1.00
wt.% C) only 17 lm. We found in 0.002 wt.% C containing ARMCOe iron (see Table 1) a thickness for the hydrogen-enriched region to be a maximum of 30 lm. Hoshihira
et al. [16] found in Al that on the surface of blisters the
hydrogen concentration is much higher than on non-blistered areas. A similar observation for Fe is shown in
Fig. 4. The physical reason for the increased hydrogen
accumulation on blister surfaces remains unclear.
4.2.2. Hydrogen on the crack surface
Another spectacular nding is the rst direct observation of hydrogen accumulations around cracks, shown
exemplarily in Fig. 6, whose existence has been mentioned
earlier, e.g. by Lynch [37]. We assume that the bright fringe
around the crack in Fig. 6 reects the scenario described by
Neeraj and Srinivasan [47] for ferritic steels. In a zone next
to the crack tip is the generation and accumulation of
excess vacancies that are stabilized by hydrogen binding
to the vacancies. The diusible lattice hydrogen will accumulate at sites of increased hydrostatic stress due to dilatation of the lattice [6]. The zone is the plasticity zone
according to the HELP theory [48], which trails the crack
path while the crack tip propagates further in the material.
Supportingly, Lo et al. [49] stated that HELP together with
the build-up of hydrogen pressure leads to blistering.
Hydrogen remains reversible trapped in the plasticity zone
[6] until, for example, a mild heat treatment (see Section
3.1) activates hydrogen eusion.
Such high local hydrogen concentration can be found
around all those cracks, which have obviously no connection to the samples surface. We can exclude that the hydrogen enrichment at the crack surface is due to chemisorption,
e.g. as binary iron hydride, because Fe has only a very weak
driving force for hydride formation [2] compared to other
elements such as V, Zr, Nb, Ta and Ti [50].
The analysis of the gas pressures inside the cracks shows
that the values are at a maximum of 20% of the yield
strength (200 MPa) of rolled non-hydrogenous ARMCO.
Such values are plausible because the tomography allows
only analyzing gas pressures in already propagated cracks
with a size >30 lm and not in smaller cracks or in pores
where we would expect higher gas pressures.
4.3. Blister formation
The origin of blisters was reported 70 years ago by Zape and Sims [51]. Recently, Ren et al. [38] and Escobar
et al. [12] observed the presence of blisters in iron as we
did, which implies that the presence of a second phase is
not a prerequisite for blister formation. Ren et al. [52]
developed a model for blistering incorporating the early
stage of hydrogen accumulation. Starting points are superabundant vacancies that aggregate together with hydrogen
to form hydrogenvacancy clusters, similar to the model of

21

Neeraj and Srinivasan [47]. In such small cavities hydrogen


atoms in the hydrogenvacancy cluster form hydrogen
molecules that can stabilize the cluster. Further hydrogen
gas accumulation increases the gas pressure. The clusters
grow through vacancies diusing into it to the size of small
pores. This is supported by our nding that pores with
diameters in the lm scale are distributed over the crack surfaces (see Fig. 9). When a certain internal gas pressure, i.e.
a critical pore size, is reached, the yield strength of the
hydrogenous material is exceeded and cracks are initiated.
Correspondingly to the increasing crack volume drops the
internal gas pressure (p V = constant, ideal gas law) and
further crack growth stops. Cracks can propagate further
with the entry and recombination of hydrogen atoms in
the cavity and subsequent increase of gas pressure past
the yield strength. If these cracks are just below the surface,
the hydrogen gas pressure in the cracks can lift up and
bulge out the exterior layer of the metal so that it resembles
a blister. Thus hydrogen induced blistering can be seen as a
special case of the plane pressure mechanism of hydrogen
embrittlement, where high pressure hydrogen forms in
microvoids near the materials surface.
We also observed cracks underneath blisters that contain no hydrogen (see Fig. 3). Here, we assume that cracks
close to the surface can have connections to the surface due
to the massive plastic deformation allowing molecular
hydrogen simply to disappear by diusion. Also Escobar
et al. [12] found cracks that initiated inside the sample
and propagated to the surface in order to release its internal gas pressure.
5. Conclusions
We observed hydrogen-assisted cracking and blistering
in electrochemically hydrogen-charged technical iron by
means of neutron tomography and classical fracture analysis with SEM and EBSD. We showed that pore formation
is a precursor for cracking with subsequent blistering. The
main conclusions are as follows:
(1) Hydrogen-assisted cracking and blistering occurred
during hydrogen charging, producing pores with
diameters in the micrometer range as well as brittle
fractured crack surfaces. The pores were located at
grain boundaries and the cracks originated from
these pores.
(2) Cracks having no connection to the samples surface
contained hydrogen gas. The gas pressure is at maximum 20% of the yield strength of non-hydrogenous
iron. The cracks had hydrogen accumulated in a plasticity zone around the cracks.
(3) Hydrogen accumulated at the samples surface with a
higher concentration at surfaces on top of blisters.
(4) The hydrogen accumulations around the cracks and
the hydrogen gas in the cracks can be removed like
diusible hydrogen by a mild temperature treatment.

22

A. Griesche et al. / Acta Materialia 78 (2014) 1422

References
[1] Oriani RA. Corrosion 1987;43:390.
[2] Dadfarnia M, Novak P, Ahn DC, Liu JB, Sofronis P, Johnson DD,
et al. Adv Mater 2010;22:1128.
[3] Akiyama E. ISIJ Int 2012;52:307.
[4] Rozenak P, Robertson IM, Birnbaum HK. Acta Metall Mater
1990;38:2031.
[5] Luppo MI, Ovejerogarcia J. Corros Sci 1991;32:1125.
[6] Olden V, Thaulow C, Johnsen R. Mater Des 2008;29:1934.
[7] Gerberich WW, Stauer DD, Sofronis P. A coexistent view of
hydrogen eects on mechanical behavior of crystals: HELP and
HEDE. In: Somerday B, Sofronis P, Jones R, editors. Eects of
hydrogen on materials. Materials Park, OH: ASM International;
2009.
[8] Mundra K, Blackburn JM, DebRoy T. Sci Technol Weld Joining
1997;2:174.
[9] Iyer RN, Pickering HW. Annu Rev Mater Sci 1990;20:299.
[10] Condon JB, Schober T. J Nucl Mater 1993;207:1.
[11] Ravi K, Ramaswamy V, Namboodhiri TKG. Mater Sci Eng A Struct
Mater 1990;129:87.
[12] Escobar DP, Minambres C, Duprez L, Verbeken K, Verhaege M.
Corros Sci 2011;53:3166.
[13] Chen J, Ai M, Wang J, Han E-H, Ke W. Corros Sci 2009;51:1197.
[14] Chung YS, Moon JH, Cho HJ, Kim HR. J Radioanal Nucl Chem
2007;272.
[15] Balden M, Lindig S, Manhard A, You J-H. J Nucl Mater
2011;414:69.
[16] Hoshihira T, Otsuka T, Tanabe T. J Nucl Mater 2009;38688:776.
[17] Terreault B. Phys Status Solidi A Appl Res 2007;204:2129.
[18] Hoshihira T, Otsuka T, Tanabe T. J Nucl Mater 2009;39091:1029.
[19] Hanada H, Otsuka T, Nakashima H, Sasaki S, Hayakawa M,
Sugisaki M. Scr Mater 2005;53:1279.
[20] Katano G, Ueyama K, Mori M. J Mater Sci 2001;36:2277.
[21] Garet M, Brass AM, Haut C, Guttierez-Solana F. Corros Sci
1998;40:1073.
[22] Revay Z, Belgya T, Szentmiklosi L, Kis Z, Wootsch A, Teschner D,
et al. Anal Chem 2008;80:6066.
[23] Banhart J. Advanced tomographic methods in materials research and
testing. Oxford: Oxford University Press; 2008.
[24] Kardjilov N, Manke I, Hilger A, Strobl M, Banhart J. Mater Today
2011;14:248.
[25] Kardjilov N, Manke I, Strobl M, Hilger A, Treimer W, Meissner M,
et al. Nat Phys 2008;4:399.

[26] Dawson M, Manke I, Kardjilov N, Hilger A, Strobl M, Banhart J.


New J Phys 2009;11.
[27] Kardjilov N, Manke I, Hilger A, Williams S, Strobl M, Woracek R,
et al. Int J Mater Res 2012;103:151.
[28] Strobl M, Grunzweig C, Hilger A, Manke I, Kardjilov N, David C,
et al. Phys Rev Lett 2008;101:123902.
[29] Manke I, Kardjilov N, Schafer R, Hilger A, Strobl M, Dawson M,
et al. Nat Commun 2010;1:125.
[30] Beyer K, Kannengiesser T, Griesche A, Schillinger B. J Mater Sci
2011;46:5171.
[31] Beyer K, Kannengiesser T, Griesche A, Schillinger B. Nucl Instrum
Methods Phys Res A 2011;651:211.
[32] Kardjilov N, Hilger A, Manke I, Strobl M, Dawson M, Williams S,
et al. Nucl Instrum Methods Phys Res A 2011;651:47.
[33] Kardjilov N, Hilger A, Manke I, Strobl M, Dawson M, Banhart J.
Nucl Instrum Methods Phys Res A 2009;605:13.
[34] Williams SH, Hilger A, Kardjilov N, Manke I, Strobl M, Douissard
PA, et al. J Instrum 2012;7:P02014.
[35] DIN EN ISO 3690. Welding and allied processes Determination of
hydrogen content in arc weld metal. 2012;28.
[36] Marrow TJ, Aindow M, Prangnell P, Strangwood M, Knott JF. Acta
Mater 1996;44:3125.
[37] Lynch S. Corros Rev 2012;30:105.
[38] Ren XC, Zhou QJ, Chu WY, Li JX, Su YJ, Qiao LJ. Chin Sci Bull
2007;52:2000.
[39] Oriani RA. Acta Metall 1970;18:147.
[40] Padhy GK, Komizo YI. Trans JWRI 2013;62:23.
[41] Takai K, Watanuki R. ISIJ Int 2003;43:520.
[42] Griesche A, Solorzano E, Beyer K, Kannengiesser T. Int J Hydrogen
Energy 2013;38:14725.
[43] Choo WY, Lee JY, Cho CG, Hwang SH. J Mater Sci 1981;16:1285.
[44] Kumnick AJ, Johnson HH. Metall Trans 1974;5:1199.
[45] Kumnick AJ, Johnson HH. Acta Metall 1980;28:33.
[46] Hadam U, Zakroczymski T. Int J Hydrogen Energy 2009;34:2449.
[47] Neeraj T, Srinivasan R, Li J. Acta Mater 2012;60:5160.
[48] Birnbaum HK, Sofronis P. Mater Sci Eng A Struct Mater
1994;176:191.
[49] Lo KH, Shek CH, Lai JKL. Mater Sci Eng R 2009;65:39.
[50] Birnbaum HK, Robertson IM, Sofronis P. Nato Adv Sci I E Appl
2000;367:367.
[51] Zape CA, Sims CE. Trans Ame Inst Min Metall Eng 1941;
145:225.
[52] Ren XC, Zhou QJ, Shan GB, Chu WY, Li JX, Su YJ, et al. Metall
Mater Trans A 2008;39A:87.

You might also like